30659
30659
https://ebookgate.com/product/introductory-biomechanics-from-
cells-to-organisms-1st-edition-c-ross-ethier/
https://ebookgate.com/product/biomechanics-and-cells-society-for-
experimental-biology-seminar-series-1st-edition-fiona-lyall/
ebookgate.com
https://ebookgate.com/product/introduction-to-bioengineering-advanced-
series-in-biomechanics-vol-2-1st-edition-y-c-fung/
ebookgate.com
https://ebookgate.com/product/human-physiology-from-cells-to-
systems-9th-edition-lauralee-sherwood/
ebookgate.com
https://ebookgate.com/product/antigen-presenting-cells-from-
mechanisms-to-drug-development-1st-edition-harald-kropshofer/
ebookgate.com
Enzymatic Fuel Cells From Fundamentals to Applications 1st
Edition Heather R. Luckarift
https://ebookgate.com/product/enzymatic-fuel-cells-from-fundamentals-
to-applications-1st-edition-heather-r-luckarift/
ebookgate.com
https://ebookgate.com/product/organizational-behavior-today-1st-
edition-stanley-c-ross/
ebookgate.com
https://ebookgate.com/product/biomaterials-from-aquatic-and-
terrestrial-organisms-1st-edition-milton-fingerman-editor/
ebookgate.com
https://ebookgate.com/product/xml-and-framemaker-1st-edition-kay-
ethier-auth/
ebookgate.com
This page intentionally left blank
Introductory Biomechanics
From Cells to Organisms
Series Editors
Professor Mark Saltzman Yale University
Professor Shu Chien University of California, San Diego
Professor William Hendee Medical College of Wisconsin
Professor Roger Kamm Massachusetts Institute of Technology
Professor Robert Malkin Duke University
Professor Alison Noble Oxford University
Cambridge University Press has no responsibility for the persistence or accuracy of urls
for external or third-party internet websites referred to in this publication, and does not
guarantee that any content on such websites is, or will remain, accurate or appropriate.
To my family, who make it all worthwhile.
c. ross eth i er
To Deborah,
and to my parents, who inspired my love of learning.
craig a. simmons
Contents
1 Introduction 1
1.1 A brief history of biomechanics 3
1.2 An outline of this book 12
References 15
2 Cellular biomechanics 18
2.1 Introduction to eukaryotic cellular architecture 18
2.2 The cell’s energy system 22
2.3 Overview of the cytoskeleton 23
2.3.1 Actin filaments 25
2.3.2 Intermediate filaments 28
2.3.3 Microtubules 28
2.4 Cell–matrix interactions 29
2.5 Methods to measure the mechanical properties of cells
and biomolecules 35
2.5.1 Atomic force microscopy 35
2.5.2 Optical trapping (“optical tweezers”) 41
2.5.3 Magnetic bead microrheometry 42
2.5.4 Micropipette aspiration 43
2.6 Models of cellular biomechanical behavior 53
2.6.1 Lumped parameter viscoelastic model of the cell 54
2.6.2 Tensegrity model of the cytoskeleton 60
2.6.3 Modeling actin filaments as a foam 69
2.6.4 Computational model of a chondrocyte in its matrix 72
2.7 Mechanotransduction: how do cells sense and respond to
mechanical events? 76
2.7.1 Mechanoreceptors 76
2.7.2 Intracellular signal transduction 78
2.7.3 Cellular response to mechanical signals 80
viii Contents
3 Hemodynamics 119
3.1 Blood rheology 119
3.1.1 Blood composition 121
3.1.2 Relationship between blood composition and rheology 124
3.1.3 Constitutive equation for blood 129
3.2 Large artery hemodynamics 130
3.2.1 Physical characteristics of blood flow patterns in vivo 130
3.2.2 Steady blood flow at low flow rates 133
3.2.3 Unsteady flow in large vessels 138
3.3 Blood flow in small vessels 142
3.3.1 Fahraeus–Lindqvist effect 143
3.3.2 “Inverse” Fahraeus–Lindqvist effect 146
3.4 Problems 147
References 161
The cover contains images that together represent the broad scope of modern
biomechanics. The figures are as follows:
r Main image: A fluorescent immunohistochemical image of an endothelial cell
isolated from the surface of a pig aortic heart valve and grown in culture. Within
the cell, the nucleus is stained blue and vimentin filaments are stained green.
Vimentin is an intermediate filament protein of the cellular cytoskeleton that
plays an important role in cellular mechanics.
r Left top: An intermediate stage from a simulation of the forced unfolding of
repeats 4 and 5 of chain A of the protein filamin. Filamin is an actin cross-linking
protein and therefore plays a role in the biomechanics of the cytoskeleton. The
simulation was based on the crystal structure of part of filamin [1], and was
carried out in NAMD [2] and visualized using the VMD package [3]. (Image
courtesy of Mr. Blake Charlebois.)
r Left middle: A sketch by the Swiss anatomist Hermann von Meyer of the
orientation of trabecular bone in the proximal human femur. This sketch was
accompanied in the original article by a sketch of the principal stress trajectories
in a crane having a shape similar to the femur. Together these sketches are
believed to have inspired “Wolff’s Law” of bone remodelling. From [4].
r Left lower: The distribution of mass transfer rates from flowing blood to cultured
vascular endothelial cells. The contoured quantity (the Sherwood number) was
computed by first measuring the topography of the endothelial cells using atomic
force microscopy and then solving the convection-diffusion equation in the
blood flowing over the cells. Mass transfer from blood to endothelial cells is
important in cell-cell signalling. (Image courtesy of Mr. Ji Zhang.)
References
r How do your bones “know” how big and strong to be so that they can support
your weight and deal with the loads imposed on them? Evidence shows that
the growth of bone is driven by mechanical stimuli [1]. More specifically,
mechanical stresses and strains induce bone cells (osteoblasts and osteoclasts)
to add or remove bone just where it is needed. Because of the obvious mechanical
function played by bone, it makes good sense to use mechanical stress as the
feedback signal for bone growth and remodeling. But biomechanics also plays
a “hidden” regulatory role in other growth processes, as the next example will
show.
r How do our arteries “know” how big to be so that they can deliver just the right
amount of blood to their distal capillary beds? There is good evidence that this
is determined in large part by the mechanical stress exerted on the artery wall by
flowing blood. Endothelial cells lining the inner arterial surface sense this shear
stress and send signals to cells deeper in the artery wall to direct the remodeling
of the artery so as to enlarge or reduce its caliber [2].
2 Introduction
r What about biomechanics in everyday life? Probably the most obvious appli-
cation of biomechanics is in locomotion (walking, running, jumping), where
our muscles generate forces that are transferred to the ground by bones and soft
connective tissue. This is so commonplace that we rarely think about it, yet the
biomechanics of locomotion is remarkably complex (watch a baby learning to
walk!) and still incompletely understood.
r Locomotion happens on many scales, from whole organisms all the way down
to individual cells. Unicellular organisms must be able to move so as to gather
nutrients, and they have evolved a variety of clever strategies to accomplish
this task [3]. In multicellular organisms, the ability of single cells to move is
essential in processes such as repair of wounds, capture of foreign pathogens,
and tissue differentiation. Force generation at the cellular level is a fascinating
topic that is the subject of much active research.
r Cells can generate forces, but just as importantly, they can sense and respond
to forces. We alluded to this above in the examples of bone remodeling and
arterial caliber adjustment, but it is not only endothelial and bone cells that
can sense forces. In fact, the ability of mechanical stress to elicit a biological
response in cells seems to be the rule rather than the exception, and some cells
are exquisitely specialized for just this task. One remarkable example is the
hair cells in the ear. These cells have bundles of thin fibers (the stereocilia)
that protrude from the apical cell surface and act as sensitive accelerometers;
as a result, the hair cells are excited by sound-induced vibrations in the inner
ear. This excitation produces electrochemical signals that are conducted by the
auditory nerve to the auditory centers in the brain in a process that we call
hearing [4,5].1
r The examples above show that biomechanics is important in homeostasis and
normal function. Unfortunately, biomechanics also plays a role in some dis-
eases. One example is glaucoma, an ocular disease that affects about 65 million
people worldwide [6]. Normally the human eye is internally pressurized, a fact
that you can verify by gently touching your eye through the closed eyelid. In
most forms of glaucoma, the pressure in the eye becomes elevated to patholog-
ical levels, and the resulting extra biomechanical load somehow damages the
optic nerve, eventually leading to blindness [7]. A second example is atheroscle-
rosis, a common arterial disease in which non-physiological stress distributions
on endothelial cells promote the disease process [8].
r What about biomechanics in the treatment of disease and dysfunction? There
are obvious roles in the design of implants that have a mechanical function,
1
Actually, the function of the hair cells is even more amazing than it first appears. The outer hair cells are active
amplifiers, changing their shape in response to mechanical stimulation and thus generating sounds. The net effect is to
apply a frequency-selective boost to incoming sounds and hence improve the sensitivity of the ear.
3 1.1 A brief history of biomechanics
such as total artificial hips [9], dental implants [10], and mechanical heart
valves [11]. In the longer term, we expect to treat many diseases by implanting
engineered replacement tissue into patients. For tissues that have a mechanical
function (e.g., heart valves, cartilage), there is now convincing evidence that
application of mechanical load to the tissue while it is being grown is essential
for proper function after implantation. For example, heart valves grown in a
bioreactor incorporating flow through the valve showed good mechanical prop-
erties and function when implanted [12]. Cartilage subjected to cyclic shearing
during growth was stiffer and could bear more load than cartilage grown with-
out mechanical stimulation [13]. We expect that biomechanics will become
increasingly important in tissue engineering, along the way leading to better
fundamental understanding of how cells respond to stresses.
The above examples should give a flavor of the important role that biomechanics
plays in health and disease. One of the central characteristics of the field is that
it is highly interdisciplinary: to be called biomechanics, there must be elements
of both mechanics and biology (or medicine). Advances in the field occur when
people can work at the frontier of these two areas, and accordingly we will try to
give both the “bio” and the “mechanics” due consideration in this book.
Another characteristic feature of biomechanics is that the topic is fairly broad. We
can get a sense of just how broad it is by looking at some of the professional societies
that fall under the heading of biomechanics. For example, in Japan alone, at least
six different professional societies cover the field of biomechanics.2 Obviously we
cannot, in a single book, go into detail in every topic area within such a broad field.
Therefore, we have given an introduction to a variety of topics, with the hope of
whetting readers’ appetites.
We can learn more about the field of biomechanics by looking at its history. In
one sense, biomechanics is a fairly young discipline, having been recognized as an
independent subject of enquiry with its own body of knowledge, societies, journals,
and conferences for only around 30–40 years. For example, the “Biomechanics and
Human Factors Division” (later to become the “Bioengineering Division”) of the
American Society of Mechanical Engineering was established in late 1966. The
International Society of Biomechanics was founded August 30, 1973; the European
2
These are the Japanese Society of Biomechanics, the Bioengineering Division of the Japan Society of Mechanical
Engineers, the Japan Society of Medical Electronics and Biological Engineering, the Association of Oromaxillofacial
Biomechanics, the Japanese Society for Clinical Biomechanics and the Japanese Society of Biorheology.
4 Introduction
Society of Biomechanics was established May 21, 1976, and the Japanese Society
of Biomechanics was founded December 1, 1984. On the other hand, people have
been interested in biomechanics for hundreds of years, although it may not have
been called “biomechanics” when they were doing it. Here we take a quick look
back through history and identify some of the real pioneers in the field. Note that
the summary below is far from exhaustive but serves to give an overview of the
history of the field; the interested reader may also refer to Chapter 1 of Fung [14]
or Chapter 1 of Mow and Huiskes [15].
Galileo Galilei (1564–1642) was a Pisan who began his university training in
medicine but quickly became attracted to mathematics and physics. Galileo was a
giant in science, who, among other accomplishments, was the first to use a tele-
scope to observe the night sky (thus making important contributions in astronomy)
and whose synthesis of observation, mathematics, and deductive reasoning firmly
established the science that we now call mechanics.3 Galileo, as part of his studies
on the mechanics of cantilevered beams, deduced some basic principles of how
bone dimensions must scale with the size of the animal. For example, he realized
that the cross-sectional dimensions of the long bones would have to increase more
quickly than the length of the bone to support the weight of a larger animal [17].
He also looked into the biomechanics of jumping, and the way in which loads are
distributed in large aquatic animals, such as whales. However, Galileo was really
only a “dabbler” in biomechanics; to meet someone who tackled the topic more
directly, we must head north and cross the English Channel.
William Harvey (1578–1657) was an English physician who made fundamen-
tal contributions to our understanding of the physiology of the cardiovascular
system, and who can be rightly thought of as one of the first biomechanicians
(Fig. 1.1). Before Harvey, the state of knowledge about the cardiovascular system
was primitive at best, being based primarily on the texts of the Roman physician
Galen (129–199?). Galen believed that the veins distributed blood to the body, while
arteries contained pneuma, a mixture of “vital spirits,” air, and a small amount of
blood. It was thought that the venous and arterial systems were not in communi-
cation except through tiny perforations in the interventricular septum separating
the two halves of the heart, so the circulatory system did not form a closed loop.
Venous blood was thought to be produced by the liver from food, after which it
flowed outward to the tissues and was then consumed as fuel by the body.4
Harvey was dissatisfied with Galen’s theories, and by a clever combination of
arguments and experimentation proved that blood must travel in a closed circuit
3
Charles Murray, in his remarkable survey of human accomplishment through the ages [16], ranks Galileo as the
second-most accomplished scientist of all time, behind (who else?) Newton.
4
It is easy to look back and ask: How could Galen have been so wrong? The answer is that he was influenced by his
predecessors; prior to Galen it was thought that arteries were filled with air and that the veins originated in the brain, for
example. The lesson to be learned: question dogma!
5 1.1 A brief history of biomechanics
Figure 1.1
Portraits of Drs. William Harvey (left) and Stephen Hales (right). Both were early biomechanicians; Harvey was a noted
English physician, while Hales was a Reverend and “amateur” scientist. Both portraits, courtesy of the Clendening
History of Medicine Library and Museum, University of Kansas Medical Center [18].
in the cardiovascular system. For example, he carried out careful dissections and
correctly noted that all the valves in veins acted to prevent flow away from the
heart, strongly suggesting that the function of the veins was to return blood to the
heart. For our purposes, his most intriguing argument was based on a simple mass
balance: Harvey reasoned that the volumetric flow of blood was far too large to be
supplied by ingestion of food. How did he do this? Using a sheep’s heart, he first
estimated the volume of blood pumped per heart beat (the stroke volume) as two
ounces of blood. Knowing the heart rate, he then computed that the heart must be
pumping more than 8600 ounces of blood per hour, which far exceeds the mass of
food any sheep would be expected to eat! In his words (italics added) [19]:
Since all things, both argument and ocular demonstration, show that the blood passes
through the lungs and heart by the force of the ventricles, and is sent for distribution to all
parts of the body, where it makes its way into the veins and porosities of the flesh, and then
flows by the veins from the circumference on every side to the center, from the lesser to the
6 Introduction
greater veins, and is by them finally discharged into the vena cava and right auricle of the
heart, and this in such a quantity or in such a flux and reflux thither by the arteries, hither
by the veins, as cannot possibly be supplied by the ingesta, and is much greater than can
be required for mere purposes of nutrition; it is absolutely necessary to conclude that the
blood in the animal body is impelled in a circle, and is in a state of ceaseless motion.
By these and additional arguments [20], Harvey deduced the closed nature of the
cardiovascular system (although he was unable to visualize the capillaries). For
our purposes, Harvey is notable because he was one of the first physicians to
use a combination of quantification, deductive reasoning, and experimentation to
understand a clinically important medical topic. Such approaches are commonplace
today but were revolutionary in Harvey’s time and even caused him to be strongly
criticized by many prominent physicians.
Giovanni Alfonso Borelli (1608–1679) is variously described as a mathemati-
cian, physicist, and physiologist, which is surely a testament to the breadth of
his interests. He worked at various universities throughout Italy, coming in con-
tact with Galileo. Notably, he spent 10 years in Pisa, where he worked with the
famous anatomist Malpighi (responsible for the discovery of the capillaries). Later
in his career, Borelli became interested in the mechanics of animal motion, and is
best known for his two-volume work on this topic, On the Movement of Animals
(De Motu Animalium), published posthumously in 1680 (Volume I) and 1681
(Volume II). In addition to the novelty of the material in these books, they are
notable for their wonderfully detailed figures illustrating biomechanical concepts
such as locomotion, lifting, and joint equilibrium (Fig. 1.2). Borelli used the princi-
ples of levers and other concepts from mechanics to analyze muscle action. He also
determined the location of the center of gravity of the human body and formulated
the theory that forward motion involved the displacement of the center of gravity
beyond the area of support and that the swinging of the limbs saved the body from
losing balance [21]. Further, he considered the motor force involved in walking and
the location of body support during walking. Borelli was also interested in respira-
tory mechanics: he calculated and measured inspired and expired air volumes. He
was able to show that inspiration is driven by muscles, while expiration is a passive
process resulting from tissue elasticity. In honor of his seminal contributions in the
field of biomechanics, the career accomplishment award of the American Society
of Biomechanics is known as the Borelli Award.
Another early biomechanician was the Reverend Stephen Hales (1677–1761),
who made contributions to both plant and animal physiology (Fig. 1.1). He is best
known for being the first to measure arterial blood pressure, now a staple of all
clinical examinations. He did this by direct arterial cannulation of his horse (in his
back yard, no less)! In his words [22,23]:
7 1.1 A brief history of biomechanics
Figure 1.2
Figure from Borelli’s classic work, De Motu Animalium (On the Movement of Animals). Panels 1–4 show how elastic
bands (representing muscles) can interact with two pivoting levers (representing bones) in a variety of geometric
configurations. Panels 5 and 6 demonstrate how the muscle and bone configurations act in humans carrying loads.
Panels 7 and 8 show various pulley arrangements, while panels 9 and 10 show how muscle action in the human arm
supports a weight R. (We will revisit this subject in Ch. 8.) The concepts may not seem advanced to modern students,
but to put things into context, it should be remembered that the first volume of De Motu Animalium was published seven
years before the appearance of Newton’s Principia.
8 Introduction
I caused a mare to be tied down alive on her back . . . having laid open the left crural
[femoral] artery about 3 inches from her belly, I inserted into it a brass pipe whose bore was
1/6 of an inch in diameter; and to that, by means of another brass pipe . . . I fixed a glass
tube of nearly the same diameter, which was 9 feet in length; then untying the ligature on
the artery, the blood rose in the tube 8 feet 3 inches perpendicular above the level of the
left ventricle of the heart . . . when it was at its full height, it would rise and fall at and after
each pulse 2, 3, or 4 inches.
Hales also improved Harvey’s estimate of cardiac stroke volume by pouring wax
at controlled pressure into the heart’s main pumping chamber (the left ventricle)
to make a casting. He then measured the volume of the wax cast by immersing it
in water, and measured its surface area by carefully covering it with small pieces
of paper covered with a measuring grid. Together with his measurements of blood
pressure, Hales then used these results to provide the first estimate of left ventric-
ular systolic (pumping) pressure, and a remarkably accurate estimate of the blood
velocity in the aorta (0.5 m/s).
Jean Léonard Marie Poiseuille (1797–1869; Fig. 1.3) was a French engineer
and physiologist who was also interested in blood flow [25]. In his thesis [27], he
described how he simplified and improved the measurement of blood pressure. His
contributions were two-fold: first, he developed the U-tube mercury manometer,
which did away with the need for an unwieldy 9 foot-long tube. Second, he used
potassium carbonate as an anticoagulant [28]. He was surprised to discover that
the pressure drop from the aorta to arteries with diameters as small as 2 mm was
negligible [29]; we know now that most of the pressure drop in the circulatory
system occurs in vessels with diameter smaller than 2 mm. Poiseuille then became
interested in laminar flow in small tubes and carried out experiments on the flow
of water through artificial glass capillaries with diameters as small as 30 μm. His
results allowed him to deduce the relationship between flow, tube dimensions, and
pressure drop, which we know today as the Hagen–Poiseuille law [30]. We will
explore the implications of this law for blood flow in Ch. 3.
Thomas Young (1773–1829) was an English physician and physicist (Fig. 1.3).
He was remarkably prodigious as a child, having learned to read “with considerable
fluency” at the age of two, and demonstrating a knack for languages, such that he
had knowledge of English, Greek, Latin, French, Italian, and Hebrew by the age
of 13 [26]. He studied medicine and practiced in London while developing and
maintaining expertise in a staggering range of areas. For example, he demonstrated
the wave theory of light, deciphered some of the first Egyptian hieroglyphics by
analysis of the Rosetta stone, helped to establish actuarial science, and lectured on
the theory of tides, surface tension, etc. In the biomedical area, he established, with
von Helmholtz, the theory of color, discovered and measured astigmatism in the
9 1.1 A brief history of biomechanics
Figure 1.3
Portraits of Drs. Jean Poiseuille (left) and Thomas Young (right). Both men did important work in physiology and
medicine, yet are familiar to engineering students: Poiseuille for his work on steady laminar incompressible flow in a
tube of uniform circular cross-section (Hagen–Poiseuille flow) and Young from his work on the elasticity of bodies
(Young’s modulus of elasticity). Poiseuille portrait reproduced with permission from [24] as modified by Sutera [25];
Young portrait by Sir Thomas Lawrence, engraved by G. R. Ward, as shown in Wood [26].
eye, and deduced that the focussing power of the eye resulted from changes in the
shape of the lens. He devised a device for measuring the size of a red blood cell, with
his measurements showing a size of 7.2 μm [26], a value that is remarkably accurate
(see Ch. 3). He also studied fluid flow in pipes and bends, and the propagation of
impulses in elastic vessels, and then applied this to analysis of blood flow in the
arteries. He correctly deduced that peristaltic motion of the artery wall did not
contribute to the circulation of blood, and instead that the motive power must
come from the heart [31]. He is most familiar to engineering students for defining
the modulus of elasticity, now known as Young’s modulus in his honor.
Julius Wolff (1836–1902) and Wilhelm Roux (1850–1924) were German physi-
cians (Fig. 1.4). Of the two, Wolff is better known to biomedical engineers because
10 Introduction
Figure 1.4
Portraits of Drs. Julius Wolff (left) and Wilhelm Roux (right). Both were German physicians who were interested in how
mechanical forces could influence the structure and development of bone. Both portraits, courtesy of the Clendening
History of Medicine Library and Museum, University of Kansas Medical Center [18].
of his formulation of “Wolff’s law” of bone remodeling. Legend has it [32] that the
structural engineer Karl Culmann saw a presentation by the anatomist Hermann
von Meyer, in which von Meyer described the internal architecture of the bone in
the head of the femur. Culmann was struck by the similarity between the pattern
of solid elements in the cancellous (“spongy”) bone of the femur and the stress
trajectories5 in a similarly shaped crane that he was designing (Fig. 1.5).6 Based
on von Meyer’s paper describing this similarity [33], as well as other data avail-
able at the time, Wolff hypothesized that bone was optimized to provide maximum
strength for a minimum mass. He then went on to formulate his “law” of bone
5
A stress trajectory is an imaginary line drawn on a surface that is everywhere tangent to the principal stress directions on
the surface. Stress trajectories help to visualize how the stress is carried by an object, and they can be used as the basis
of a graphical procedure for determining stress distributions in bodies. This graphical solution method is now obsolete,
having been replaced by computational methods.
6
This certainly emphasizes the importance of interdisciplinary interaction in biomedical engineering!
11 1.1 A brief history of biomechanics
Figure 1.5
Comparison of internal architecture of cancellous bone in the head of a femur (large drawing at right) and the stress
trajectories in the head of a crane (large drawing at left). The smaller drawings provide details of the mechanics of the
crane and the stress distributions in various structures. This picture originally appeared in the article by von Meyer [33],
as reproduced in [32].
remodeling [34,35]:7 “Every change in the form and the function of a bone or of
their function alone is followed by certain definite changes in their internal archi-
tecture, and equally definite secondary alterations in their external confirmation,
in accordance with mathematical laws.” In simpler terms, Wolff stated that bone
will adapt its internal architecture in response to external constraints and loads.
Wolff went on to claim a rigorous similarity between cancellous bone architecture
and stress trajectories. Cowin has shown that this is based on a false comparison of
apples and oranges (i.e., a continuous material vs. a porous one) [32] and argued
persuasively that Wolff gets rather too much credit for his remodeling law, possi-
bly because he was a more prolific writer on this topic. Cowin [32] suggested that
the anatomist Roux should get at least as much credit as Wolff for this “law” of
7
Wolff’s original book was in German [34], but an English translation exists [35].
12 Introduction
remodeling, and we are inclined to agree. Roux was very interested in developmen-
tal biology and physiology, carrying out his doctoral work on the factors governing
the bifurcation of blood vessels [36]. He was convinced that mechanical and phys-
ical principles played important roles in development (e.g., [37]) and carried out
seminal experimental studies in this area. In 1894, he also founded a journal enti-
tled Archiv für Entwicklungsmechanik (Archives of Developmental Mechanics) and
served as its editor for many years, stirring up his fair share of controversy along the
way [38]. He may be rightly thought of as the first developmental biomechanician.
In the years since Wolff and Roux, there have been many advances in biome-
chanics. This is not the place to try to provide a complete history of biomechanics,
and we will not list all of the outstanding investigators who have worked (or are
working) in this fast-growing field. However, it is worth mentioning one other
investigator. An important event in the maturation of the field of biomechanics
was the publication, in 1981, of the book Biomechanics: Mechanical Properties
of Living Tissues, by Yuan-Cheng Fung (Fig. 1.6). Fung was born in 1919 and
was trained as an aeronautical engineer, a field in which he made many techni-
cal contributions in the early years of his career. However, in the late 1950s, he
became interested in biomechanics and consequently changed his research focus
away from aeronautics. In addition to his 1981 book, Fung has also written sev-
eral other books in biomechanics [40,41]. Y.-C. Fung is generally regarded as the
“father” of modern biomechanics and has the rare distinction of being a member
of the US National Academy of Engineering, the US Institute of Medicine, and
the US National Academy of Sciences. Membership in all three of these learned
societies is surely a testament to the abilities of Dr. Fung, but it is also a reflection of
the highly interdisciplinary nature of biomechanics, which (when done properly)
should tightly integrate engineering, medicine, and biology.
Figure 1.6
Portrait of Y.-C. Fung, who has played an important role in the establishment of biomechanics as a modern, rigorous
discipline, primarily through the publication of an influential series of books on biomechanics. Reproduced with
permission from [39].
cellular biomechanics. How stiff is a cell, and how do we measure this parameter?
How does the cell anchor itself to substrates and to neighbors? How does the
cell respond to external forces, and what implications does this have for tissue
organization?
At a higher level, physiologists subdivide the body into organs, which are tissues
specialized for a specific purpose, and systems, which are collections of organs
working in concert. In this book, we will also consider the biomechanics of some
of the body’s systems. There are many such systems; here is a partial list and
description of their functions, with emphasis on the biomechanical aspects.
Circulatory system. This system delivers nutrients and picks up waste products
from the cells, as well as delivering signaling molecules, such as hormones, between
different organs. The flow of blood will be studied in Ch. 3; other biomechanical
aspects of the circulatory system are discussed in Ch. 4.
14 Introduction
Lymphatic system. Excess fluid is passively collected from tissues and returned
to the heart via a network of ducts and channels that make up the lymphatic system.
The lymphatic system is also an important locus for immune function. We will
not examine the lymphatic system in detail, although it is briefly touched upon
in Ch. 5.
Nervous and sensory systems. The nervous system consists of the nerves and
brain and is responsible for signaling and control within the body. Its operation
is highly complex and will not be considered in this book. The sensory organs
provide input to the nervous system; we will briefly consider ocular biomechanics
in Ch. 6.
Respiratory system. In order to oxidize foodstuffs, O2 must be delivered to the
blood and CO2 must be removed from it. This is accomplished by exposing the
blood to the air through a very thin membrane of enormous surface area. This
membrane is convoluted and folded to form a large number of small sacs within
the lung. The respiratory system consists of the lungs, plus structures that assist
air passage in and out of the lungs. It will be considered in Ch. 7.
Urinary system. The urinary system consists of the kidneys, ureters, urinary
bladder, and urethra. The kidneys are responsible for removing waste products
from blood and for the production of urine, which is then stored in the bladder and
excreted through the urethra. The biomechanics of the urinary system is fascinating
and complex [42–45]. Unfortunately, consideration of the urinary system is beyond
the scope of this book.
Muscular system. Muscles are specialized tissues that generate force upon
appropriate stimulation. Muscular action is required for locomotion (movement
of the body), motion of individual body parts, and bulk transport of materials
within the body (e.g., pumping of blood by the heart). Muscles, and how they
effect motion, will be discussed in Chs. 8 and 10.
Skeletal system. This framework of bones and soft connective tissues (cartilage,
ligaments, and tendons) provides a rigid, supportive, and protective structure for the
body. The bony skeleton also provides attachments for muscles, serves as a system
of levers for movement and locomotion, and has important metabolic functions.
Bones, cartilage, ligaments, and tendons will be discussed in greater detail in
Ch. 9.
Digestive system. The digestive system comprises the gastrointestinal tract
(mouth to anus) plus the liver, gall bladder, and pancreas. The digestive system
is responsible for ingestion and breakdown of food, delivery of foodstuffs to the
blood, and waste excretion. We will not consider the digestive system in this book.
Immune system. This system consists of specialized cells and molecules dis-
tributed throughout the body (in organs such as the spleen and as cells in the blood-
stream and interstitial fluid). It is responsible for identification and destruction of
15 References
foreign entities, including viruses and bacteria. Consideration of the immune sys-
tem is beyond the scope of this book.
References
8. R. Ross. The pathogenesis of atherosclerosis: a perspective for the 1990s. Nature, 362
(1993), 801–809.
9. J. P. Paul. Strength requirements for internal and external prostheses. Journal of
Biomechanics, 32 (1999), 381–393.
10. J. T. Steigenga, K. F. al Shammari, F. H. Nociti, C. E. Misch and H. L. Wang. Dental
implant design and its relationship to long-term implant success. Implant Dentistry,
12 (2003), 306–317.
11. Q. Yuan, L. Xu, B. K. Ngoi, T. J. Yeo and N. H. Hwang. Dynamic impact stress
analysis of a bileaflet mechanical heart valve. Journal of Heart Valve Disease, 12
(2003), 102–109.
12. S. P. Hoerstrup, R. Sodian, S. Daebritz, J. Wang, E. A. Bacha, D. P. Martin et al.
Functional living trileaflet heart valves grown in vitro. Circulation, 102 (2000),
III44–III49.
13. S. D. Waldman, C. G. Spiteri, M. D. Grynpas, R. M. Pilliar and R. A. Kandel.
Long-term intermittent shear deformation improves the quality of cartilaginous tissue
formed in vitro. Journal of Orthopaedic Research, 21 (2003), 590–596.
14. Y. C. Fung. Biomechanics: Mechanical Properties of Living Tissues, 2nd edn (New
York: Springer Verlag, 1993).
15. V. C. Mow and R. Huiskes (ed.) Basic Orthopaedic Biomechanics and Mechano-
Biology, 3rd edn (Philadelphia, PA: Lippincott Williams & Wilkins, 2005).
16. C. Murray. Human Accomplishment: The Pursuit of Excellence in the Arts and
Sciences, 800 BC to 1950 (New York: HarperCollins, 2003).
17. A. Ascenzi. Biomechanics and Galileo Galilei. Journal of Biomechanics, 26 (1993),
95–100.
18. Clendening History of Medicine Library and Museum. Clendening Library Portrait
Collection. Available at http://clendening.kumc.edu/dc/pc/ (2005).
19. W. Harvey. Exercitatio anatomica de motu cordis et sanguinis in animalibus. [An
Anatomical Study of the Motion of the Heart and of the Blood in Animals.]
(1628).
20. W. C. Harrison. Dr. William Harvey and the Discovery of Circulation (New York:
MacMillan, 1967).
21. A. J. Thurston. Giovanni Borelli and the study of human movement: an historical
review. ANZ Journal of Surgery, 69 (1999), 276–288.
22. S. Hales. Statical Essays, Containing Haemastaticks. [With an Introduction by Andre
Cournand.] (New York: Hafner, 1964).
23. S. Hales. Foundations of anesthesiology. Journal of Clinical Monitoring and
Computing, 16 (2000), 45–47.
24. M. Brillouin. Jean Leonard Marie Poiseuille. Journal of Rheology, 1 (1930), 345–348.
25. S. P. Sutera and R. Skalak. The history of Poiseuille law. Annual Review of Fluid
Mechanics, 25 (1993), 1–19.
26. A. Wood. Thomas Young, Natural Philosopher 1773–1829 (New York: Cambridge
University Press, 1954).
17 References
The cell is the building block of higher organisms. Individual cells themselves
are highly complex living entities.1 There are two general cell types: eukaryotic
cells, found in higher organisms such as mammals, and prokaryotic cells, such as
bacteria. In this chapter, we will examine the biomechanics of eukaryotic cells only.
We will begin by briefly reviewing some of the key components of a eukaryotic cell.
Readers unfamiliar with this material may wish to do some background reading
(e.g., from Alberts et al. [1] or Lodish et al. [2]).
Walls (the membranes). These barriers are primarily made up of lipids in a bilayer
arrangement, augmented by specialized proteins. They serve to enclose the cell,
the nucleus, and individual organelles (with the exception of the cytoskeleton,
which is distributed throughout the cell). The function of membranes is to create
compartments whose internal materials can be segregated from their surroundings.
For example, the cell membrane allows the cell’s interior to remain at optimum
levels of pH, ionic conditions, etc., despite variations in the environment outside
the cell. The importance of the cell membrane is shown by the fact that cell death
almost invariably ensues if the cell membrane is ruptured to allow extracellular
materials into the cell.
A framework (the cytoskeleton). This organelle consists of long rod-shaped
molecules attached to one another and to other organelles by connecting molecules.
1
What defines a living system? A living system must satisfy the following five characteristics:
r it must show complex organization (specialization)
r it must be able to metabolize (assimilate “food,” transform it, and excrete it)
r it must show responsiveness, including the ability to adapt to differing conditions
r it must be able to reproduce
r it must have evolutionary capability.
Individual cells exhibit each of these characteristics.
19 2.1 Introduction to eukaryotic cellular architecture
The cytoskeleton gives the cell form, allows it to move, helps to anchor the cell to its
substrate and neighbors, and speeds the transport of materials within certain types
of cells. We will consider the cytoskeleton in greater detail later in this chapter.
Engines (the mitochondria). These organelles produce most of the basic energy-
containing molecules from certain substrates such as glucose. Then these energy-
containing molecules are used by other subsystems within the cell.
A command center (the nucleus). The genetic material which codes for
molecules synthesized by the cell is mostly contained in the nucleus, although
there is a small amount of mitochondrial DNA. The synthesis of proteins and other
important biomolecules is initiated by transcription of genetic information coded
in DNA into messenger RNA (mRNA). These mRNA coding sequences leave the
nucleus where they are eventually turned into proteins by . . .
Factories (the endoplasmic reticulum). These production centers synthesize
biomolecules needed by the cell. They take their “orders” from the nucleus in
the form of mRNA.
Packaging plants (the Golgi apparatus). Proteins produced by the endoplasmic
reticulum are not “ready for prime time.” They typically must undergo a series
of folding steps and post-translational modifications before they have biologi-
cal activity. This very important task is handled by the Golgi apparatus, which
takes the protein output of the endoplasmic reticulum and trims, modifies, and
packages it in membrane-delimeted structures (the vesicles) that are sent to var-
ious locations within or outside the cell. Misfolded and otherwise defective pro-
teins must be disposed of immediately since they are potentially harmful to the
cell.
A disposal system (the lysozomes). This system of vesicles contains enzymes
(catalytic proteins) which act to break down metabolic by-products, misfolded
proteins, ingested extracellular material, and other unwanted substances.
Microfilaments
Plasma membrane
Secretory vesicle
Microtubule
Lysosome
Centrioles
Peroxisome
Free ribosome
Nucleus
Bound ribosome
Nucleolus
Granular
endoplasmic
reticulum Nuclear envelope
Nuclear pore
Agranular
endoplasmic
reticulum
Figure 2.1
Structures and organelles found in most human cells. This diagram is highly schematized but serves to indicate the
major features of the cellular organelles. From Vander et al. [3]. Reproduced with kind permission of the McGraw-Hill
Companies.
Figure 2.2
Transmission electron micrograph of an insulin-producing pancreatic cell, showing several of the structures depicted
schematically in Fig. 2.1. A prominent nucleus delimited by the nuclear envelope (membrane) is present, as are several
organelles in the cytoplasm: mitochondria, Golgi complex, and endoplasmic reticulum with associated ribosomes. A
second cell is visible at the top left of the image. Sample stained with uranyl acetate and lead citrate. Micrograph
courtesy of Mr. Steven Doyle, University of Toronto.
One of the remarkable things about most cells is how good they are at sens-
ing relatively small levels of mechanical stimulation, while living in a constantly
changing biomechanical environment. How do they do this? Many details of this
process, known as mechanotransduction, are unknown, but in Section 2.7 we will
discuss, in general terms, current thinking on how adherent cells are able to sense
and respond to mechanical stimulation.
22 Cellular biomechanics
How does the cell utilize food energy? When we eat a meal, the constituent food-
stuffs are acted upon by the digestive enzymes and broken down into simpler
compounds, transferred into the bloodstream across the intestinal walls, and then
transported throughout the body.
Individual cells are therefore presented with a complex mixture of compounds
from which they must obtain energy. The cell solves this problem by having spe-
cialized “energy plants” (mitochondria), which are able to use compounds such
as glucose and fatty acids to produce a common energy-containing molecule that
all cellular organelles can use. This common molecule is adenosine triphosphate
(ATP), formed from adenosine diphosphate (ADP) and phosphate (PO2− 3 ) in the
following reaction:
+
ADP + PO2−
3 + energy + 2H → ATP
Energy is stored in the chemical bond between ADP and PO2− 3 (Fig. 2.3). A mechan-
ical analogue is a spring, which starts in an uncompressed state (ADP) and is then
compressed and held in place by a catch (PO2− 3 ). The organelles can “release the
catch” to produce energy, with by-products ADP and PO2− 3 . We say, therefore, that
ATP is the common currency of energy within the cell.
It is important to note that ADP and phosphate are recycled, to be once more
combined in the mitochondria to yield ATP. This is similar to the movement of
primary loop cooling water in a thermal generating station, and it emphasizes that
ATP is merely a transient carrier of energy within the cell (Fig. 2.4).
23 2.3 Overview of the cytoskeleton
NH2
N C C
N adenine
CH
C CH
N N O O O
O CH2 O P O P O P O− + HOH
C H H C O− O− O−
H C C
OH OH
Ribose
ATP
NH2
N C C
N
CH
CH
N C N O O O
O CH2 O P O P O− + HO P O− + H+ + 7 kcal/mole
C H H C O− O− O−
H C C H Inorganic phosphate
OH OH
ADP
Figure 2.3
Structure of ADP and ATP. The breakdown of ATP to ADP and inorganic phosphate yields 7 kcal/mole energy. From
Vander et al. [3]. Reproduced with kind permission of the McGraw-Hill Companies.
Water + steam To
Fuel + turbines
oxygen
Return
from
CO2 +
Water turbines
H2O
Energy Flow
Carbohydrates, Organelle
ATP
protein, fats + products
oxygen
Mitochondria Organelles
CO2 + Raw
H2O ADP + PO32− materials
Figure 2.4
Energy flow in a thermal generating station (top) and a cell (bottom). Note that primary loop cooling water is analogous
to ATP in that it is a transient vehicle for energy storage which is recycled.
The cytoskeleton consists of three types of filament, each with a specialized pro-
tein composition: actin filaments (7–9 nm in diameter), intermediate filaments
(10 nm in diameter), and microtubules (approximately 24 nm in diameter). Actin
filaments are also called microfilaments or – in skeletal muscle cells – thin fila-
ments. The interaction between all three filament types helps to determine the cell’s
mechanical behavior. We will briefly review the function of each of these filament
types.
25 2.3 Overview of the cytoskeleton
Figure 2.5
Scanning electron micrograph of the actin component of the cytoskeleton within a rat fibroblast adhering to an
N-cadherin-coated glass cover slip (A). A high-magnification inset of the boxed region is shown in (B). The rich, highly
interconnected actin network is clearly visible. Cells were extracted with a detergent solution, fixed in glutaraldehyde,
post-fixed, and gold sputter coated before visualization. Images courtesy of Dr. Tarek El Sayegh, Faculty of Dentistry,
University of Toronto.
Actin exists within the cell in two forms, as a globular protein (G-actin) and as
a filamentous protein (F-actin). G-actin has a molecular weight of approximately
43 kDa, and consists of a single polypeptide chain. Monomeric G-actin binds one
Ca2+ and one molecule of ATP. F-actin is formed by the polymerization of G-actin,
which causes the bound ATP to be hydrolyzed to ADP and a phosphate ion (Fig. 2.7,
color plate). The ADP remains bound to the actin subunit within the F-actin chain.
F-actin chains are dynamic structures that grow and break down according to
their position within the cell and the activities of the cell at any given instant. F-actin
filaments are polarized, having an end where G-actin monomers are preferentially
added (the fast-growing “barbed” or “+” end) and an end where the filament is
either slowly growing or disassembled (the “pointed” or “–” end). Thus, individ-
ual actin monomers move along filaments, tending to be added at the + end and
26 Cellular biomechanics
Figure 2.6
Cytoskeleton of the brush border of intestinal epithelial cells. Tight actin filaments are evident in the microvilli, the
finger-like structures in the top half of the image. The actin extends from microvilli into the cytoplasm of the cell, where it
connects with a network of actin, intermediate filaments, myosin, and other cytoskeletal proteins. The scale bar in the
lower left corner is 0.1 μm. Reproduced with permission from Bershadsky and Vasiliev [4] and from the Journal of Cell
Biology, 1982, 94, 425–443 by copyright permission of the Rockefeller University Press.
27 2.3 Overview of the cytoskeleton
moving to the – end, in a process known as treadmilling (Fig. 2.8, color plate).
The polymerization and breakdown of F-actin are regulated by several proteins,
including actin depolymerizing factor/cofilin, members of the gelsolin/villin pro-
tein family, and CapZ. The lifetime of actin filaments, the length of the filaments,
the percentage of actin in polymeric form, and the number of barbed ends change
as a function of the cell’s activity. For example, in confluent bovine aortic endothe-
lial cells, the mean filament lifetime is approximately 40 min and about 70% of
the cell’s total actin is present as F-actin [8]. Confluent cells tend to be relatively
quiescent, exhibiting only modest amounts of cellular movement. In contrast, in
subconfluent endothelial cells, which are more active, the mean filament lifetime
is only approximately 8 min and only approximately 40% of the cell’s actin is
present in polymerized form [8].
F-actin exists within the cell in several different forms. In all cells, it is present
in a thin layer adjacent to the cell membrane, in the so-called cortical actin layer.
This layer helps to anchor transmembrane proteins to cytoplasmic proteins (see
Section 2.4) and generally provides mechanical strength to the cell. It also anchors
the centrosomes at opposite ends of the cell during mitosis, and helps the cell to split
into two parts during cytokinesis. In many cells, actin is also present in long bundles
that criss-cross the cell, known as stress fibers. Stress fibers presumably reinforce
the cortical actin layer and are also important for the transport of organelles and
cellular locomotion. Finally, in skeletal muscle cells, actin is highly abundant
and interacts with myosin filaments (thick filaments) to create the actin–myosin
complex, responsible for muscle contraction.
Because of its biomechanical importance, a number of authors have experimen-
tally estimated the Young’s modulus of F-actin filaments. Some of these measure-
ments are summarized in Table 2.1, which shows a fairly tight range of values of
order 2 GPa, approximately 100 times less than the modulus for steel.
Another Random Scribd Document
with Unrelated Content
ff. ) that besides the twelve, Jesus chose other seventy also, and
sent them two and two before him into all the districts which he
intended to visit on his last journey, that they might proclaim the
approach of the kingdom of heaven. As the other Evangelists have
no allusion to this event, the most recent critics have not hesitated
to make their silence on this head a reproach to them, particularly to
the first Evangelist, in his supposed character of apostle. 69 But the
disfavour towards Matthew on this score ought to be moderated by
the consideration, that neither in the other gospels, nor in the Acts,
nor in any apostolic epistle, is there any trace of the seventy
disciples, who could scarcely have passed thus unnoticed, had their
mission been as fruitful in consequences, as it is commonly
supposed. It is said, however, that the importance of this
appointment lay in its significance, rather than in its effects. As the
number of the twelve apostles, by its relation to that of the tribes of
Israel, shadowed forth the destination of Jesus for the Jewish
people; so the seventy, or as some authorities have it, the seventy-
two disciples, were representatives of the seventy or seventy-two
peoples, with as many different tongues, which, according to the
Jewish and early Christian view, formed the sum of the earth’s
inhabitants, 70 and hence they denoted the universal destination of
Jesus and his kingdom. 71 Moreover, seventy was a sacred number
with the Jewish nation; Moses deputed seventy elders (Num. xi.
16 , 25 ); the Sanhedrim had seventy members; 72 the Old
Testament, seventy translators.
Had Jesus, then, under the pressing circumstances that mark his
public career, nothing more important to do than to cast about for
significant numbers, and to surround himself with inner and outer
circles of disciples, regulated by these mystic measures? or rather, is
not this constant preference for sacred numbers, this assiduous
development of an idea to which the number of the apostles
furnished the suggestion, wholly in the spirit of the primitive
[333]Christian legend? This, supposing it imbued with Jewish
prepossessions, would infer, that as Jesus had respect to the twelve
tribes in fixing the number of his apostles, he would extend the
parallel by appointing seventy subordinate disciples, corresponding
to the seventy elders; or, supposing the legend animated by the
more universal sentiments of Paul, it could not escape the
persuasion that to the symbol of the relation of his office to the
Israelitish people, Jesus would annex another, significative of its
destination for all the kindreds of the earth. However agreeable this
class of seventy disciples may have always been to the church, as a
series of niches for the reception of men who, without belonging to
the twelve, were yet of importance to her, as Mark, Luke and
Matthew; we are compelled to pronounce the decision of our most
recent critic precipitate, and to admit that the Gospel of Luke, by its
acceptance of such a narrative, destitute as it is of all historical
confirmation, and of any other apparent source than dogmatical
interests, is placed in disadvantageous comparison with that of
Matthew. We gather, indeed, from Acts i. 21 f. that Jesus had more
than the twelve as his constant companions; but that these formed a
body of exactly seventy, or that that number was selected from
them, does not seem adequately warranted 73. [334]
§ 76.
The identity of the discourses being established, the first effort was
to conciliate or to explain the divergencies between the two
accounts so as to leave their credibility unimpeached. In reference to
the different designation of the locality, Paulus insists on the ἐπὶ of
Luke, which he interprets to imply that Jesus stood over the plain,
and therefore on a hill. Tholuck, more happily, distinguishes the level
space, τόπος πεδινὸς, from the plain properly so called, and regards
it as a less abrupt part of the mountain. But as one Evangelist makes
Jesus ascend the mountain to deliver his discourse, while the other
makes him descend for the same purpose, these conciliators ought
to admit, with Olshausen, that if Jesus taught in the plain, according
to Luke, Matthew has overlooked the descent that preceded the
discourse; or if, as Matthew says, Jesus taught seated on the
mountain, Luke has forgotten to mention that after he had
descended, the pressure of the crowd induced him to reascend
before he commenced his harangue. And without doubt each was
ignorant of what he omits, but each knew that tradition associated
this discourse with a sojourn of Jesus on a mountain. Matthew
thought the mountain a convenient elevation for one addressing a
multitude; Luke, on the contrary, imagined a descent necessary for
the purpose: hence the double discrepancy, for he who teaches from
a mountain is sufficiently elevated over his hearers to sit, but he who
teaches in a plain will naturally stand. The chronological
divergencies, as well as the local, must be admitted, if we would
abstain from fruitless efforts at conciliation. 3
ebookgate.com